• Aucun résultat trouvé

Superconductivity of NbSe3

N/A
N/A
Protected

Academic year: 2021

Partager "Superconductivity of NbSe3"

Copied!
8
0
0

Texte intégral

(1)

HAL Id: jpa-00209337

https://hal.archives-ouvertes.fr/jpa-00209337

Submitted on 1 Jan 1981

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Superconductivity of NbSe3

A. Briggs, P. Monceau, M. Nunez-Regueiro, M. Ribault, J. Richard

To cite this version:

A. Briggs, P. Monceau, M. Nunez-Regueiro, M. Ribault, J. Richard. Superconductivity of NbSe3.

Journal de Physique, 1981, 42 (10), pp.1453-1459. �10.1051/jphys:0198100420100145300�. �jpa-

00209337�

(2)

Superconductivity of NbSe3

A. Briggs, P. Monceau, M. Nunez-Regueiro, M. Ribault (*) and J. Richard

Centre de Recherches sur les Très Basses Températures, C.N.R.S., B.P. 166 X, 38042 Grenoble Cedex, France

(Reçu le 6 février 1981, révisé le 18 mai, accepté le 10 juin 1981)

Résumé.

2014

Nous avons mesuré les transitions résistives supraconductrices de NbSe3 sous champ magnétique

pour des pressions supérieures à celles nécessaires à la suppression de l’onde de densité de charge qui apparaît

à 59 K à pression ambiante. NbSe3 est alors un supraconducteur massif anisotrope avec un effet Meissner total.

Les anisotropies des champs magnétiques critiques sont bien décrites par un modèle de masses effectives qui sont

en bon accord avec celles déduites des mesures de conductivité ou des mesures d’oscillations Shubnikov-de Haas.

A pression ambiante, nous avons mesuré des transitions avec résistivité nulle pour des échantillons de NbSe3 compactés mais sans effet Meissner total. Cette supraconductivité n’est pas massive et est due vraisemblablement

aux barrières entre les domaines formant le cristal.

Abstract.

2014

We report measurements of the resistive superconducting transitions of NbSe3 as a function of

magnetic field for sufficient pressures to destroy the lower CDW. These indicate that NbSe3 is a bulk anisotropic superconductor with a total Meissner effect. The critical field anisotropies are well described by an effective mass

model and the values obtained are in good agreement with those obtained by conductivity or Shubnikov-de Haas measurements. At ambient pressure we have measured zero-resistivity transitions for compacted samples, but do

not find a total Meissner effect. Hence this superconductivity is not a bulk effect but is probably associated with barriers between domains in the sample.

Classification

Physics Abstracts

62.50

-

74.10

-

74.60

Introduction.

-

Among the family of metal-transi- tion trichalcogenides intensively studied these last years, only TaSe3 and NbSe3 [1] remain metallic at

helium temperature [2, 3]. TaSe3 has a superconducting

transition at 2.2 K [4]. NbSe3 undergoes two inde- pendent charge-density wave transitions with asso-

ciated lattice distortions at 145 K (Tl) and

59 K (T2) [2, 5]. Two other resistive anomalies have been observed : the resistivity drops between 2.2 K

and 1.5 K and it is constant down to 0.4 K where it decreases again. But down to 7 mK no zero-resistivity

has been measured. Magnetization measurements indicate that a very small part of the NbSe3 sample

becomes diamagnetic below 2.2 K where the resistivity drops [3]. A filamentary model of superconductivity

has been proposed to explain this low temperature behaviour [6]. Under pressure the CDW transition temperatures decrease. It has been shown that

dT 2/dP = - 6.25 K/kbar [7]. For this pressure and above NbSe3 is a bulk superconductor with a complete

Meissner effect. Zero-resistivity has been measured

on doped NbSe3 samples with impurities of tanta-

lum [8], titanium [9] or zirconium [10], for tempe- (*) Permanent address : Laboratoire de Physique des Solides, Université Paris-Sud, Bât. 510, 91405 Orsay, France.

LE

JOURNAL DE PHYSIQUE

-

T. 42, 10, OCTOBRE 1981

ratures around 1.5 K but no magnetization on these doped samples have yet been reported. We have

observed that a zero-resistivity transition occurs in

compacted pure NbSe3 samples. The critical tempe-

rature is around 0.9 K.

Hereafter we first review the resistive and magnetic

measurements on pure NbSe3 filaments. Then we

describe the experiments on compacted NbSe3 samples

with different size grains. Finally we report measure-

ments on some NbSe3 monocrystals under pressure.

We have determined the critical magnetic fields parallel and perpendicular to the chain direction.

NbSe3 under pressure is an anisotropic super- conductor which near Tc can be described with an

effective mass model. The effective masses deduced from the critical magnetic fields for the different

crystallographic orientations are in good agreement with those obtained by conductivity and Shubnikov- de Haas measurements.

1. Pure NbSe3.

-

The drop in resistivity below

2.2 K is sample dependent [3]. This drop can be a

few percent of the resistivity above 2.5 K but we have

measured 80 % on a bundle of filaments. Fleming

does not observe such a drop on very pure samples

with resistance ratios of 200 [11]. Below 2.2 K the

94

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198100420100145300

(3)

1454

resistivity is strongly dependent on the current density

and the critical current decreases exponentially with

temperature. The transition can be suppressed with

current densities higher than 1 A/cm2. Previous

d.c. initial susceptibility measurements on a NbSe3 sample of 300 mg at 50 mK in a magnetic field of

1 Oe indicated that less than 10- 3 of the volume could be superconducting [3]. More accurate magne- tization measurements using SQUID magnetometers have been recently reported [12]. On NbSe3 samples loosely assembled Buhrman et al. show that a small

diamagnetic contribution appears around 2 K where the resistivity drops and that at the lowest temperature (a few tens of mK) the diamagnetic signal is 1 % of

-

V/4 n in agreement with the limit of our previous

measurements. They described a model where indi- vidual planes become superconducting around 2 K.

At lower temperature the coupling between planes

would increase leading to a 2D-3D superconducting

transition.

The temperature dependence of the superconducting

critical current and the absence of a total Meissner effect exclude a bulk superconductivity and we proposed a filamentary model for the superconducting

transition at 2.2 K [6]. However NbSe3 is very different

morphologically from, for instance (SN)x. This poly-

mer is an assembly of fibers of around 100 A diameter which dominate the superconducting properties [13].

Very recently Fung and Steeds [14] have observed

that the CDW lattice consisted of elongated domains

with a typical size of 2 g x 200 A x 200 A for a pure

NbSe3 sample. Taking into account this observation

we have assumed that the CDW order parameter is

zero inside the borders between two domains in which the phase of the CDW is constant [15]. At low temperature these borders become superconducting

which lead to a multiconnexç superconductivity.

This assumption can account for the sample dependent

behaviour because of the different repartition of

domains in different samples.

2. Compacted NbSe3.

-

In compounds like NbSe3

where the easy path for electrons is along the chains,

the effect of impurities consists of breaking chains ; consequently the transport properties are principally

due to these interrupted strands. The resistivity

measurements between 300 K and 4.2 K that we have made on samples doped with titanium, zirconium,

tantalum or after electron irradiation [16] show a strong increase of the resistivity at low temperature.

The same behaviour has been observed with proton irradiation [9]. Another method of « dirtying » the samples is to use powders with different grain sizes.

We have crushed NbSe3 filaments to obtain powder.

This powder was passed through several sieves with

grain size less than 100 g (sample F6) and with grain

size between 43 03BC and 60 g (sample F8). The compac- tion of the powder was done under an hydraulic press at 100 bars at room temperature. The sample dimen-

sions are typically 10 x 1.5 x 1 mm3. The d.c. resis- tance is measured by a classical four lead contacts.

The current density is typically 2 x 10- 2 A/cm2. The

insert of figure 1 shows the resistivity variation bet-

ween 300 K and 4.2 K [17]. The resistivity variation

can be separated in two parts : a semiconducting-like

variation all the more important as the grain size is

smaller and the resistivity variation of the pure NbSe3.

The resistivities at 4.2 K normalized to the room tem-

perature resistivity are respectively 2.07 for F6 and

3.54 for F8. It can be seen that anomalies due to the two CDWs at 145 K and 59 K remain which ensures

that, although no X-rays determination have been

performed on the powder, we measure the real NbSe3 phase.

Fig. 1.

-

Low temperature resistance of compacted powder of NbSe3. The grain size of sample F6 is less than 100 Il and for sample

F8 between 40 u and 63 Il. In the insert, the resistance variation normalized to room temperature as a function of temperature for F6, F8 and a monocrystal of NbSe3.

Around 2 K the resistivity saturates. We have measured the resistivity of F6 and F8 in a 3He refri- gerator. The resistivity variation is shown in figure 1.

The resistivity drops to zero for the two samples. If

we define the critical temperature at the middle of the

transition, Te is 0.8 K for F6 and 0.9 K for F8. Initial

susceptibility measurements made on F6 show a small

diamagnetism of the order of 6 % down to 200 mK

and increase of this diamagnetism around 100 mK up to 25 % of the total diamagnetism.

In figure 2 we show the variation of the resistivity

of F6 as a function of magnetic field at different tem-

peratures. Due to the random orientation of each

(4)

Fig. 2.

-

Variation of the resistance of F6 normalized to its room

temperature resistance as a function of magnetic field at different temperatures. In the insert the normalized resistance at 21 kG

as a function of temperature. The resistance increases monotonically

when the superconducting state is suppressed by the magnetic field.

grain and the anisotropy of the crystal, the critical

fields are not very well defined. But at sufficiently high magnetic fields it can be seen that the magneto- resistance follows the same variation which is the normal state one. We can obtain the resistivity in a

fixed magnetic field (21 kG) where the sample is in

the normal state (except for the two lowest tempera-

tures where the values are obtained by extrapolation).

The insert of figure 2 indicates that when the super-

conducting state is suppressed (by the magnetic field)

the normal resistivity increases monotonically below

2 K.

Our compacted samples are pure NbSe3 but disor-

dered in the sense that many barriers have been intro- duced. Although the two CDWs are présent, a zero-

resistivity is observed below 0.9 K without a complete

Meissner effect. To ascertain that zero-resistivity was

not a surface property of the compacted sample we

cut F6 longitudinally in three parts and we remeasured separately these three parts. No difference in behaviour from the original specimen was observed.

Zero resistivity has also been measured on doped NbSe3 samples. The resistivity below 2.5 K is current dependent for samples doped with 0.1 % of titanium [9]

and with zirconium [10] although the effect of current is less important than for pure NbSe3 [3]. A super- conductivity transition was reported for a NbSe3

with 5 % of tantalum at 1.5 K [8]. It was stated that the superconductivity arose because the lower CDW

was smeared out as similar to the pressure effect. If this is true, as in the case of pressure, a complete

Meissner effect must be observed. However no magne- tization measurements have yet been performed on doped samples.

Lee and Rice [18] have calculated that the size of the domains where the phase of the CDW is constant decreases with the amount of impurities. So in the

doped samples as in our compacted samples we expect that there is a great number of normal barriers around CDW domains and therefore a superconduct- ing percolation path at low temperature leading to

a zero-resistance. However there is not a total Meissner effect.

3. NbSe3 under pressure.

-

We have measured the

resistivity of two single crystals of NbSe3 under pres-

sure below 4.2 K. The typical dimensions were

5 x 0.02 x 0.005 mm. Precession photographs taken

with filtered MoK radiation ensured that the crystals

were single crystals with the plane (b, c) in the plane

of the ribbon. The crystals were mounted on an araldic

disc in a beryllium-copper chester clamp capable of retaining pressures up to 11 kbar at 300 K resulting

in about 7.5 kbar at 4.2 K. Pressures were measured at room temperature and nitrogen temperature with

a pressure cycled manganin resistance placed near

the specimen using the calibration data of Itske- vich [19]. For a fixed room temperature pressure, the

manganin gauge gave a reproducible nitrogen tem- perature resistance. The fluid used was an isopentane- methyl 2-pentane mixture. The samples were those

used for Shubnikov-de Haas oscillations measure- ments that we reported previously [7]. For such

measurements only two probes on the crystals were

necessary which excluded absolute d.c. resistance measurements.

NbSe3 becomes superconducting when the lower CDW transition is totally suppressed by pressure. In the insert of figure 3 we show the pressure dependence

of the CDW transition T2 and the resistive super-

conducting transition. Above 5.5 kbar, NbSe3 is

superconducting with a sharp transition and Tc

decreases when the pressure is increased. In the critical pressure range where T2 varies sharply with pressure,

T, decreases rapidly and a superconducting transition

cannot be detected below 5 kbar. The continuous variation of the resistance for pressures in this range could be due to pressure inhomogeneities. More hydrostatic measurements with helium gas are under- way to study this possible coexistence between super-

conductivity and CDW state.

Figure 3 shows the resistive transitions for pressures of 5.5 kbar and 7.2 kbar. At 7.2 kbar, NbSe3 undergoes

a very sharp superconducting transition at 3.4 K with

a width of 0.1 K. The residual resistance of 4 Q is the contact resistance because of the two probes

measurement which varies between 3.2-4 Q for the different pressures. At 5.5 kbar there is a break in the resistance variation at 3.5 K followed by a continuous

decrease of the resistance down to 1.5 K. But if we

extrapolate the variation of the magnetic critical

fields at H

=

0 as we report below, we find a critical temperature of 3.5 ± 0.1 K. Unfortunately because of

the two contacts measurements we were unable to measure at ambient pressure the drop in resistivity

at 2.2 K.

(5)

1456

Fig. 3.

-

Resistive transitions of a monocrystal of NbSe3 for

pressures of 5.5 and 7.2 kbar. The measurements were made with two probes and the residual 3-4 Q is the contact resistance. In the insert the variation of the lower CDW transition and the super-

conducting transition as a function of pressure. When the CDW is suppressed NbSe3 becomes a bulk superconductor.

Fig. 4.

-

Magnetization of a NbSe3 sample formed with many fibers at 5.4 kbar and 1.37 K.

We have previously reported initial susceptibility

measurements of a specimen consisting of many fibers

(specimen C weight of 0.56 g in reference [7]) under

pressure. The specimen showed a complete Meissner

effect above 5 kbar at 2.5 K. The difference between the superconducting critical temperatures obtained by

resistive and magnetic measurements is not well understood and we are trying to clarify this point.

Figure 4 shows the magnetization of sample C at

1.37 K for a pressure of 5.4 kbar. This magnetization

variation is typical of a superconductor with strong flux trapping.

For magnetic measurements we need to know the

angular orientations. The sample has the chains

parallel to b and its large face is parallel to (b, c).

For the transverse orientation (H 1 b) we define the

azimutal angle 9 between H and c : 9

=

0 corresponds

to H parallel to c and ç

=

n/2 at H parallel at a*

(which is different from a for the monoclinic structure).

The polar angle 0 is defined in the plane (a*, b) and

0=0 corresponds to H//b. The resistive transitions in magnetic fields were performed in two different configurations : firstly with 0

=

n/2 and 9

=

0 at 7.7 kbar, for one sample only, in the course of our

Shubnikov-de Haas oscillations measurements [7] and secondly at 5.5 and 7.4 kbar with the pressure cell inserted in an electromagnet capable of rotating between 0=0 and 0

=

n/2 with 9

=

n/2, for the two samples. We have drawn in figure 5 typical resistive

transitions as a function of magnetic field at the

pressures of 5.5 and 7.4 kbar for the orientations ç = n/2, 0=0 and 9

=

n/2, 0

=

n/2. In figure 6 we

have drawn the temperature variation of the critical fields for the three orientations investigated for the

same sample. The extrapolation of H 1. (0

=

n/2, ç

=

0)

at 7.7 kbar gives T,

=

3.25 K compared to Tc

=

3.4 K

at 7.4 kbar and 3.5 K at 5.5 kbar. This result confirms the decrease of the critical temperature for pressures

greater than those necessary for suppressing the lower

Fig. 5.

-

Resistive transitions of a monocrystal of NbSe3 for

pressures of 5.5 and 7.4 kbar as a function of magnetic field. The orientation of the magnetic field with the crystallographic axis are

shown on the insert. ç is the azimutal angle with H in the plane perpendicular to the chain or b axis, 0 is the polar angle with H rotating in the plane (a, b). The resistive transitions correspond

to 9

=

n/2, 0

=

n/2 (H 1 b) and ç

=

n/2, 0

=

0 (H//b).

(6)

Fig. 6.

-

Variation of the critical magnetic fields as a function of the temperature for three different orientations. At 7.7 kbar for 0 = n/2, w

=

0. At 5.5 and 7.2 kbar for (p

=

n/2, 0

=

n/2 and for

cp

=

n/2, 0=0.

-

CDW. The extrapolation towards zero of the critical field H1 (0

=

n/2, ç

=

n/2) and H Il (0

=

0, ç

=

n/2) gives the same critical temperatures as those obtained in zero field. We observe a slight upward curvature

in the temperature variation of H Il for the two pres-

sures and for Hl (0

=

n/2, ç

=

n/2) at 5.5 kbar. Such behaviour has already been reported in two-dimen- sional dichalcogenides and is characteristic of aniso-

tropic superconductors [20]. In table 1 we give the slopes dH.l,ll/dT near T,, for the three pressures.

Critical magnetic fields have also been measured in tantalum doped samples but without studying the anisotropy in the transverse orientation [8]. Thus we

can only compare thé dH)j/dTB

The critical fields in a multilayered superconducting

material can be calculated using the model of Law-

rence and Doniach of Josephson coupled planes [21].

When H is perpendicular to the planes, the fluxoids have a circular cross section and the vortex current circulates freely as for a bulk superconductor. In this

case, the critical field HC2J.

=

qJo/2 nçn where

qJo

=

hcl2 e is the flux quantum, ç the coherence

length. When H is parallel to the layers the vortex

current circulates inside a layer and tunnels from one

layer to another as a Josephson current. Near T, the

coherence length is much larger than the layer sepa- ration and an effective mass model can be used. The

fluxoids are ellipsoidals and the critical field is defined

by :

01,

Following the same model the angular dependence of He2 between the parallel and perpendicular orienta-

tions is [22] :

At lower temperature the coherence length may become lower than the layers separation and the nor-

mal vortex cores are located between the layers.

HC2U largely exceeds the bulk value. For T T*

the layers decouple and the critical field diverges.

This cross over temperature between a 3D-2D beha- viour has been calculated by Klemm et al. [23] and

Boccara et al. [24]. The decoupling between the layers explains the upward curvature of He 2 measured experimentally.

This theory of Josephson junctions has been extend- ed to filamentary superconductors by Turkevich and

Klemm [25]. They consider an assembly of filaments oriented in the Z direction which forms a lattice with dimensions a and b. When H is applied perpen- dicular to the chains, the vortex current follows the chain and tunnels by the Josephson effect from one chain to another as when H is parallel to the planes

in a layered superconductor. When H is parallel to

the chains the vortex currents are entirely Josephson

tunnel currents and the system behaves as a Josephson grid. Near T, the coherence lengths are much larger

than the chains separation a and b and a generalized

effective mass model can be described. In the per-

pendicular orientation the conductivity is no longer isotropic and the angular azimutal dependence of H C2 becomes :

Table 1.

-

Initial slopes of the critical magnetic fields dH/dT in kG/K.

NbSe3

NbSe3

(7)

1458

mx and rny being the effective masses respectively in

the x and y directions.

The polar dependence of H e2 when H is rotated from

being perpendicular to parallel to the chains is :

At low temperatures the vortex core can fit between the chains. Below T T * the chains decouple leading

to the divergence of the critical magnetic field but T * is not the same when H is parallel or perpendicular

to the chains.

Such a model can be applied to NbSe3. The structure

can be seen as infinite planes of chains two atoms

thick separated by Van der Waals gaps parallel to

the (b, c) plane. In the transverse orientation (0

=

n/2)

at 7.4 kbar the extrapolation at T

=

0 of Hc, gives

a coherence length ç(O)

=

360 A much larger than the

chain separation. Consequently the effective mass

model near T, is applicable to NbSe3. We find that Hl is greater for qJ

=

0 than for ç

=

n/2 indicating

a lower effective mass in this direction. The ratio between H eJ. (9 = 0, 0 = n/2) and Hc (qJ

=

n/2, 0

=

n/2)

is equal to 2 which corresponds to an effective mass

ratio of 4. Anisotropy in the transverse orientation has been measured by magnetoresistance and Shub-

nikov-de Haas oscillation studies. A ratio of 3 bet-

ween the cyclotron effective masses in the same

orientations (ç

=

0, 0

=

n/2) and

=

n/2, 0

=

n/2)

has been measured [26] which is in a relatively good

agreement with the value obtained from the aniso- tropy of the superconducting critical fields.

We show in figure 7 the polar dependence of the

critical field as a function of 0. We have drawn in the

same figure the theoretical variation given by equa- tion (4) with the transverse critical field equals to

1.15 k0e. The best fit is obtained with Gx

=

0.1820

which is in excellent agreement with the value of

Fig. 7.

-

Angular dependence of the critical magnetic field at

7.2 kbar as a function of the polar angle 0. The curve is the theoretical variation (Eq. (4)).

0.1849 obtained by the ratio of the critical fields with 9

=

n/2, 0

=

n/2 and ç

=

0, 0

=

n/2. This model

gives an anisotropy of 30 between the effective masses

when the electrons are travelling parallel to b or parallel to c. This result is in good agreement with the anisotropy of conductivity measured by Ong and

Brill [27].

The upward curvature of Hie Il indicates the strong anisotropy of NbSe3 and some decoupling between

chains at low temperature in agreement with a

Josephson coupled chains description. It must be

noted that the analysis of the critical fields under pressure shows that NbSe3 is three-dimensional near

T, with some decoupling between chains at lower temperatures. This is the opposite of the statement

of Buhrman et al. [12] where they expect that individual

planes to become superconducting around 2 K and these planes to couple together at lower temperatures

to form a 3D superconductor.

4. Conclusions.

-

We have studied the low tem-

perature properties of NbSe3. The behaviour under pressure is completely different from that at ambient pressure when we apply sufficient pressure to destroy

the lower CDW transition (T2

=

59 K at P

=

0).

NbSe3 is a bulk superconductor with a total Meissner effect. The total diamagnetism has been observed with initial susceptibility measurements and the

magnetization curves are typical of an irreversible type II superconductor with strong flux trapping. Due

to the morphology of NbSe3 with a large anisotropy

in the orientation perpendicular to the chains, the

effective mass model of Josephson coupled chains is

reasonable near T,,. From the anisotropy of the critical fields we find a ratio of effective masses in the plane perpendicular to b of 4 and 30 in the plane (a*, b)

which is in good agreement with values obtained from Shubnikov-de Haas oscillations and conducti-

vity anisotropies.

At ambient pressure, pure NbSe3 shows a drop in resistivity at 2.2 K and a very small diamagnetism is

measured at very low temperature. Powdered samples

show a zero resistance below 0.9 K. Doped samples

show also zero resistance but no total Meissner effect has yet been reported. We consider that, unlike NbSe3

under pressure, this superconductivity is not a bulk

effect but due to barriers between domains in the

sample.

Acknowledgments.

-

We would like to thank A. Meerschaut, P. Molinié, and J. Rouxel from the Laboratoire de Chimie Minérale de Nantes for pro-

viding us with the samples and M. Papoular and

M. Renard for helpful discussions.

(8)

References

[1] MEERSCHAUT, A. and ROUXEL, J., J. Less Common Metals 39

(1975) 197.

[2] CHAUSSY, J., HAEN, P., LASJAUNIAS, J. C., MONCEAU, P., WAYSAND, G., WAINTAL, A., MEERSCHAUT, A., MOLI-

NIE, P. and ROUXEL, J., Solid State Commun. 20 (1976) 759.

[3] HAEN, P., LAPIERRE, F., MONCEAU, P., NUNEZ-REGUEIRO, M.

and RICHARD, J., Solid State Commun. 26 (1978) 725.

[4] SAMBONGI, T., YAMAMOTO, M., TSUTSUMI, K., SHIOZAKI, Y., YAMAYA, K. and ABE, Y., J. Phys. Soc. Japan 42 (1977)

1421.

[5] FLEMING, R. M., MONCTON, D. E. and MCWHAN, D. B., Phys. Rev. B 18 (1978) 5560.

[6] HAEN, P., MIGNOT, J. M., MONCEAU, P. and NUNEZ-RFGUEIRO, M., J. Physique Colloq. 39 (1978) C6-703.

[7] BRIGGS, A., MONCEAU, P., NUNEZ-REGUEIRO, M., PEYRARD, J., RIBAULT, M. and RICHARD, J., J. Phys. C 13 (1980) 2117.

[8] FULLER, W. W., CHAIKIN, P. M. and ONG, N. P., Solid State Commun. 30 (1979) 689.

[9] FULLER, W. W., Thesis, University of California at Los Angeles (1980), unpublished and

FULLER, W. W., CHAIKIN, P. M. and ONG, N. P., Solid State Commun. 39 (1981) 547.

[10] NISHIDA, K., SAMBONGI, T. and IDO, M., J. Phys. Soc. Japan 48 (1980) 331.

[11] FLEMING, R. M., Phys. Rev. B 22 (1980) 5606.

[12] BUHRMAN, R. A., BASTUSCHECK, C. M., SCOTT, J. C. and KULICK, J. D., in Inhomogeneous Superconductors, edited by D. W. Gubser, T. L. Francavilla, S. A. Wolf and J. R. Leibowitz (New York, American Institute of Phy- sics) 1980.

[13] AZEVEDO, L. J., CLARK, W. G., DEUTSCHER, G., GREENE, R. L., STREET, G. B. and SUTER, L. J., Solid State Commun. 19

(1976) 197.

[14] FUNG, K. K. and STEEDS, J. W., Phys. Rev. Lett. 45 (1980) 1696.

[15] MONCEAU, P., RICHARD, J. and RENARD, M., Phys. Rev., in press.

[16] MONCEAU, P., RICHARD, J. and LAGNIER, J. Phys. C 14 (1981) 2995.

[17] NUNEZ-REGUEIRO, M., Thesis, Université de Grenoble (1979), unpublished.

[18] LEE, P. A. and RICE, T. M., Phys. Rev. B 19 (1979) 3970.

[19] ITSKEVICH, E. S., Pribary Tekhn. Eksp. 4 (1962) 148, Cryogenics

4 (1964) 365.

[20] WOOLLAM, J. A., SOMOANO, R. B. and O’CONNOR, P., Phys.

Rev. Lett. 32 (1974) 712.

[21] LAWRENCE, W. E. and DONIACH, S., Proceedings of the 12th

International Conf. on Low Temperature Physics, edited by Eizo Kanda (Academic Press of Japan, Kyoto) 1971,

p. 361.

[22] MORRIS, R. C., COLEMAN, R. V. and BHANDARI, R., Phys. Rev.

B 5 (1972) 895.

[23] KLEMM, R. A., LUTHER, A. and BEASLEY, M. R., Phys. Rev.

B 12 (1975) 877.

[24] BOCCARA, N., CARTON, J. P. and SARMA, G., Phys. Lett. 49A (1974) 165.

[25] TURKEVICH, L. A. and KLEMM, R. A., Phys. Rev. B 19 (1979)

2520.

[26] MONCEAU, P., Solid State Commun. 24 (1977) 331.

[27] ONG, N. P. and BRILL, J. W., Phys. Rev. B 18 (1978) 5265.

Références

Documents relatifs

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

61 See Supplemental Material for additional electron diffraction images, crystallographic details inferred from single-crystal X-ray diffraction, additional images of

Abstract.- We have been synthesizing high Tc systems based on NbsGe in which another transition metal atom, either Mo, Ti or Zr, is partially substituted for Nb.. Systems have

Les premiers donnent lieu & une diminution ou & une augmentation de la tedpQature critique, selon l'ampleur de l'anisotropie du (c gap n d'knergie ou

These we suggest form spin polarons with the Cu 3d moments, which combine into bipolarons which are the pairs needed for

The present results on a-LaZn, together with those obtained with other amorphous La-based alloys and a-PdZr alloy, indicate an inverse relation between the residual

The exportent of the dielectric constant divergence of a mixture of an insulator and a metal is identical to that of the electrical conductivity in the case of superconductor

• un large savoir-faire en formation initiale, continue et en double-compétence avec leur programme commun phare, le MAE, destiné avant tout aux non-gestionnaires • un maillage